This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Some Solitons on Homogeneous Almost α\alpha-Cosymplectic 33-Manifolds and Harmonic Manifolds

Naeem Ahmad Pundeer Department of Mathematics
Jadavpur University
Kolkata-700032, India.
[email protected]
   Paritosh Ghosh Department of Mathematics
Jadavpur University
Kolkata-700032, India.
[email protected]
   Hemangi Madhusudan Shah Harish-Chandra Research Institute
A CI of Homi Bhabha National Institute
Chhatnag Road, Jhunsi, Prayagraj-211019, India.
[email protected]
   Arindam Bhattacharyya Department of Mathematics
Jadavpur University
Kolkata-700032, India
[email protected]
Abstract.

In this paper, we investigate the nature of Einstein solitons, whether it is steady, shrinking or expanding on almost α\alpha-cosymplectic 33-manifolds. We also prove that a simply connected homogeneous almost α\alpha-cosymplectic 33-manifold, admitting a contact Einstein soliton, is an unimodular semidirect product Lie group. Finally, we show that a harmonic manifold admits a non-trivial Ricci soliton if and only if it is flat. Thus we show that rank one symmetric spaces of compact as well as non-compact type are stable under a Ricci soliton. In particular, we obtain a strengthening of Theorem 11 and Theorem 22 of [1].

Key words and phrases:
Almost α\alpha-cosymplectic manifold, Harmonic manifold, Ricci soliton, Einstein soliton
1991 Mathematics Subject Classification:
53B40, 58B20, 53C25, 53D15

1. Introduction

The study of solitons, in particular Ricci solitons, on Riemannian manifolds play a vital role in understanding the geometry of underlying manifold. It is very interesting to study Ricci and Einstein solitons on almost α\alpha-cosymplectic 33-manifolds. Recently, Jin and Ximin [10] showed that a simply connected homogeneous almost α\alpha-cosymplectic 33-manifold, admitting contact Ricci solitons, is cosymplectic; and the manifold under consideration is an unimodular semidirect product Lie group 2A\mathbb{R}^{2}\rtimes_{A}\mathbb{R}, where A=(0bb0)A=\left(\begin{array}[]{cc}0&b\\ -b&0\end{array}\right), equipped with a flat left invariant cosymplectic structure.

Motivated by this result we show in this paper that, if a simply connected homogeneous almost α\alpha-cosymplectic 33-manifold, with some additional hypothesis, admits a contact Einstein soliton, then the manifold is an unimodular semidirect product Lie group GG of type G0bb¯=2AG_{0b\overline{b}}=\mathbb{R}^{2}\rtimes_{A}\mathbb{R}, where A=(0bb0)0A=\left(\begin{array}[]{cc}0&b\\ -b&0\end{array}\right)\neq 0. And also GG is the Lie group E~2\tilde{E}^{2} equipped with its flat left invariant cosymplectic structure (see Corrollary 3.5). In order to prove this result, we first obtain a characterization of almost α\alpha-cosymplectic 33-manifold admitting contact Einstein solitons, which is the main theorem (Theorem 3.4) of Section 3. To establish this aforementioned theorem we derive an identity (Lemma 3.3) involving scalar curvature, Lie derivative of the metric and Ricci operator on a Riemannian manifold admitting Einstein soliton. We also give some conditions on α\alpha for contact Einstein solitons to be steady, shrinking or expanding on almost α\alpha-cosymplectic 33-manifolds (see Theorem 3.1).

Another interesting topic in the differential geometry is the geometry of harmonic manifolds. In this paper, we prove that a harmonic manifold admits a non-trivial Ricci soliton if and only if it is flat. The flat harmonic manifold admits Ricci solitons of steady, expanding or shrinking type. We also determine the corresponding potential function. In fact, Busemann function on n\mathbb{R}^{n} turns to be the potential function in case of steady solitons (see Theorem 4.1 of Section 4).
Note that any rank one symmetric space is harmonic. Therefore, in particular, we obtain that there are no non-trivial Ricci solitons on rank one symmetric spaces. Thus harmonic manifolds, in particular, rank one symmetric spaces are stable under a Ricci soliton. It is shown in Theorem 11 and Theorem 22 of [1] that any small perturbation of the non-compact symmetric metric is flown back to the original metric under an appropriately rescaled Ricci flow. Thus we obtain the strengthening of this result in case of non-compact rank one symmetric spaces. Moreover, we also obtain that compact rank one symmetric spaces are stable under the Ricci solitons.

The paper is divided into four sections. Section 2 is devoted to the preliminaries about Ricci soliton, Einstein soliton, almost α\alpha-cosymplectic 33-manifolds and harmonic manifolds. In Section 3, we prove our main results on almost α\alpha-cosymplectic 33-manifold admitting contact Einstein solitons, as stated above. In the last section, we prove the main result about harmonic manifolds admitting Ricci solitons.

2. Preliminaries

In this section, we discuss some notions required to prove the results of this paper.

2.1. Ricci solitons

Ricci solitons are the self similar solutions of the Ricci flow. The concept of Ricci flow was first introduced by Hamilton [8] in (1982), motivated by the work of Eells and Sampson [7] on harmonic map and the flow was given by the equation

gt=2S,\frac{\partial g}{\partial t}=-2S,

where SS is the Ricci tensor.
Ricci solitons are the generalizations of the Einstein metrics and are the solutions of the equation

Ric(g)+12𝔏Xg=λg,Ric(g)+\frac{1}{2}\mathfrak{L}_{X}g=\lambda g, (1)

where Ric(X,Y)=S(X,Y)Ric(X,Y)=S(X,Y) is the Ricci curvature tensor, 𝔏X\mathfrak{L}_{X} is the Lie derivative along the direction of the vector field XX and λ\lambda is a real constant. The soliton is said to be shrinking if λ>0\lambda>0, steady if λ=0\lambda=0 and expanding if λ<0\lambda<0.
Tashiro [17] proved very important result for complete Einstein manifolds admitting Ricci solitons.

Theorem 2.1.

[17] Let (M,g)(M,g) be a complete Riemannian nn-manifold admitting a nontrivial function ff such that Hessf=λg\operatorname{Hess}f=\lambda g, then (M,g)(M,g) is isometric to a complete warped product metric and must have one of the three forms:

  1. (1)

    M=×N,g=dr2+ρ2(r)gNM=\mathbb{R}\times N,~{}~{}g=dr^{2}+\rho^{2}(r)g_{N},

  2. (2)

    M=n,g=dr2+ρ2(r)dsn12,r0M=\mathbb{R}^{n},~{}~{}g=dr^{2}+\rho^{2}(r)ds_{n-1}^{2},~{}~{}r\geq 0,

  3. (3)

    M=Sn,g=dr2+ρ2(r)dsn12,r[a,b]M=S^{n},~{}~{}g=dr^{2}+\rho^{2}(r)ds_{n-1}^{2},~{}~{}r\in[a,b].

2.2. Einstein solitons

The Einstein solitons are the generalization of the Ricci solitons, was first introduced by Catino and Mazzieri [4] in (2016). They are the solutions of the equation

𝔏Vg+2S=(2λ+r)g,\mathfrak{L}_{V}g+2S=(2\lambda+r)g, (2)

where, Ricci tensor S(X,Y)=g(X,QY)S(X,Y)=g(X,QY), QQ being the Ricci operator, rr is the scalar curvature, λ\lambda\in\mathbb{R} is a constant and VV is known as potential vector field.
Einstein solitons are the self-similar solutions of the Einstein flow,

tg+2S=rg.\frac{\partial}{\partial t}g+2S=rg.

It is said to be steady if λ=0\lambda=0, shrinking if λ>0\lambda>0 and expanding if λ<0\lambda<0.

2.3. Almost contact metric manifolds

In order to define contact metric manifolds, we need the concept of Reeb vector field.
Reeb vector field [3]: A global vector field ξ\xi on a contact manifold M2n+1M^{2n+1}, equipped with a global 11-form η\eta, is called Reeb vector field or characteristic vector field, if any vector field XX satisfies η(ξ)=1\eta(\xi)=1 and dη(X,ξ)=0d\eta(X,\xi)=0.

Almost contact manifold [3]: Let MM be a Riemannian manifold of dimension (2n+1)(2n+1), n1n\geq 1. M2n+1M^{2n+1} is said to have an almost contact structure (φ,ξ,η)(\varphi,\xi,\eta), if there exists a (1,1)(1,1)-tensor φ\varphi, a global vector field ξ\xi and a 11-form η\eta such that

φ2X=X+η(X)ξ,η(ξ)=1,\displaystyle\varphi^{2}X=-X+\eta(X)\xi,~{}~{}~{}~{}\eta(\xi)=1, (3)

for any vector field XX on MM, where ξ\xi is the Reeb vector field. The manifold MM equipped with the structure (φ,ξ,η)(\varphi,\xi,\eta) is called an almost contact manifold.

Almost contact metric manifold [3]: A Riemannian metric gg is said to be compatible with an almost contact structure (φ,ξ,η)(\varphi,\xi,\eta), if

g(φX,φY)=g(X,Y)η(X)η(Y),g(\varphi X,\varphi Y)=g(X,Y)-\eta(X)\eta(Y), (4)

holds for any X,Yχ(M)X,Y\in\chi(M) and (M,φ,ξ,η,g)(M,\varphi,\xi,\eta,g) is called an almost contact metric manifold.

Normal almost contact metric manifold [3]: An almost contact metric manifold is said to be normal, if for any X,Yχ(M)X,Y\in\chi(M) the tensor field N=[φ,φ]+2dηξN=[\varphi,\varphi]+2d\eta\otimes\xi vanishes everywhere on the manifold, where [φ,φ][\varphi,\varphi] is the Nijenhuis tensor of φ\varphi.

Homogeneous almost contact metric manifold [10]: An almost contact metric manifold (M,φ,ξ,η,g)(M,\varphi,\xi,\eta,g) is said to be homogeneous, if there exists a connected Lie group GG of isometries acting transitively on MM leaving η\eta invariant.

2.4. Cosymplectic manifolds

A (2n+1)(2n+1)-dimensional manifold is said to be a cosymplectic manifold [11], if it admits a closed, 1-form η\eta and 22-form Φ\Phi such that ηΦn\eta\wedge\Phi^{n} is a volume element, where Φ(X,Y)=g(φX,Y)\Phi(X,Y)=g(\varphi X,Y) is a 22-form on M2n+1M^{2n+1}.

Almost cosymplectic manifold [11]: If η\eta and Φ\Phi are not closed but ηΦn\eta\wedge\Phi^{n} is a volume form, then the manifold is called almost cosymplectic manifold.

α\alpha-cosymplectic manifold [14]: An almost cosymplectic manifold is said to be α\alpha-cosymplectic if dη=0d\eta=0 and dΦ=2αηΦd\Phi=2\alpha\eta\wedge\Phi for some constant α\alpha.

Almost α\alpha-cosymplectic manifold [11]: An almost α\alpha-cosymplectic manifold is defined as an almost contact metric manifold with dη=0d\eta=0 and dΦ=2αηΦd\Phi=2\alpha\eta\wedge\Phi, for any constant α\alpha. In particular, the almost α\alpha-cosymplectic manifold is

  • almost α\alpha-Kenmotsu if α0\alpha\neq 0,

  • almost cosymplectic if α=0\alpha=0,

  • almost Kenmotsu if α=1\alpha=1.

Harmonic vector field [16]: A characteristic vector field ξ\xi on an almost α\alpha-cosymplectic manifold is harmonic if and only if ξ\xi is an eigenvector field of the Ricci operator QQ.

2.5. Almost α\alpha-cosymplectic 33-manifold

In this article, we will mainly focus on 33-dimensional almost α\alpha-cosymplectic manifold. In what follows, we will be using the following results.

Theorem 2.2.

[14] An almost α\alpha-cosymplectic 33-manifold is α\alpha-cosymplectic if and only if 𝔏ξh=0\mathfrak{L}_{\xi}h=0, where h=12𝔏ξφh=\frac{1}{2}\mathfrak{L}_{\xi}\varphi.

Any almost α\alpha-cosymplectic 33-manifold satisfies important relationships between Φ,ξ\Phi,\xi and hh.

Lemma 2.3.

[14] Let M2n+1M^{2n+1} be an almost α\alpha-cosymplectic 33-manifold, then we have,

ξφ=0,ξ=0,hφ+φh=0,hξ=0,\displaystyle\nabla_{\xi}\varphi=0,~{}~{}~{}~{}\nabla\xi=0,~{}~{}~{}~{}h\varphi+\varphi h=0,~{}~{}~{}~{}h\xi=0, (5)

with

Xξ=αφ2XφhX.\nabla_{X}\xi=-\alpha\varphi^{2}X-\varphi hX. (6)

We would require some identities on the φ\varphi-bases [3] and the following table of the Levi-Civita connection.

Proposition 2.4.

[14] On almost α\alpha-cosymplectic 33-manifold, there exists φ\varphi-bases satisfying

he=σe,hφe=σφe,hξ=0,\displaystyle he=\sigma e,~{}~{}~{}~{}h\varphi e=-\sigma\varphi e,~{}~{}~{}~{}h\xi=0,

with σ\sigma a local smooth eigen-function of hh.

Theorem 2.5.

[14] The Levi-Civita connection on almost α\alpha-cosymplectic 33-manifold are given by,

{ee=aφeαξ,φee=bφe+σξ,ξe=μφe,eφe=ae+σξ,φeφe=beαξ,ξφe=μe,eξ=αeσφe,φeξ=σe+αφe,ξξ=0,\begin{cases}&\nabla_{e}e=-a\varphi e-\alpha\xi,~{}~{}~{}~{}\nabla_{\varphi e}e=-b\varphi e+\sigma\xi,~{}~{}~{}~{}\nabla_{\xi}e=\mu\varphi e,\\ &\nabla_{e}\varphi e=ae+\sigma\xi,~{}~{}~{}~{}\nabla_{\varphi e}\varphi e=be-\alpha\xi,~{}~{}~{}~{}\nabla_{\xi}\varphi e=-\mu e,\\ &\nabla_{e}\xi=\alpha e-\sigma\varphi e,~{}~{}~{}~{}\nabla_{\varphi e}\xi=-\sigma e+\alpha\varphi e,~{}~{}~{}~{}\nabla_{\xi}\xi=0,\end{cases} (7)

where a=g(eφe,e)a=g(\nabla_{e}\varphi e,e), b=g(φee,φe)b=-g(\nabla_{\varphi e}e,\varphi e) and μ=g(ξe,φe)\mu=g(\nabla_{\xi}e,\varphi e) are smooth functions.

The Ricci operator on almost α\alpha-cosymplectic 33-manifold is known explicitly [14].

Proposition 2.6.

[14] The Ricci operator QQ on almost α\alpha-cosymplectic 33-manifold is given by,

{Qξ=(2α2+trh2)ξ+(2bσe(σ))φe(2aσ+(φe)(σ))e,Qφe=(2bσe(σ))ξ+(α2+r2+trh22+2σμ)φe+(ξ(σ)+2ασ)e,Qe=(2aσ+(φe)(σ))ξ+(ξ(σ)+2ασ)φe+(α2+r2+trh222σμ)e.\begin{cases}&Q\xi=-(2\alpha^{2}+\operatorname{tr}h^{2})\xi+(2b\sigma-e(\sigma))\varphi e-(2a\sigma+(\varphi e)(\sigma))e,\\ &Q\varphi e=(2b\sigma-e(\sigma))\xi+(\alpha^{2}+\frac{r}{2}+\frac{\operatorname{tr}h^{2}}{2}+2\sigma\mu)\varphi e+(\xi(\sigma)+2\alpha\sigma)e,\\ &Qe=-(2a\sigma+(\varphi e)(\sigma))\xi+(\xi(\sigma)+2\alpha\sigma)\varphi e+(\alpha^{2}+\frac{r}{2}+\frac{\operatorname{tr}h^{2}}{2}-2\sigma\mu)e.\\ \end{cases} (8)

Furthermore, the scalar curvature r=trQr=\operatorname{tr}Q is given by

r=6α2trh22(a2+b2)2(φe)(a)+2e(b).r=-6\alpha^{2}-\operatorname{tr}h^{2}-2(a^{2}+b^{2})-2(\varphi e)(a)+2e(b). (9)

The structure of simply-connected, homogeneous almost α\alpha-cosymplectic 33-manifold, admitting a contact Ricci soliton, is very well known.

Theorem 2.7.

[10] Let MM be a simply-connected, homogeneous almost α\alpha-cosymplectic 33-manifold admitting a contact Ricci soliton. Then MM is an unimodular semidirect product Lie group GG of type G0bb¯=2AG_{0b\overline{b}}=\mathbb{R}^{2}\rtimes_{A}\mathbb{R}, where A=(0bb0)A=\left(\begin{array}[]{cc}0&b\\ -b&0\end{array}\right), equipped with a flat left invariant cosymplectic structure. Moreover, we have the following:

  1. (1)

    If A=0A=0, i.e., b=0b=0, GG is the abelian Lie group 3\mathbb{R}^{3} equipped with its flat left invariant cosymplectic structure.

  2. (2)

    If A0A\neq 0, i.e., b0b\neq 0, GG is the Lie group E~2\tilde{E}^{2} equipped with its flat left invariant cosymplectic structure.

2.6. Harmonic manifolds

A complete Riemannian manifold (Mn,g)(M^{n},g) is said to be harmonic, if for any pMp\in M, the volume density ωp(q)=det(gij(q))\omega_{p}(q)=\sqrt{\det(g_{ij}(q))} in normal coordinates, centered at any pMp\in M is a radial function [2]. Thus,

Θ(r)=rn1det(gij(q)),\Theta(r)=r^{n-1}\sqrt{\det(g_{ij}(q))},

is density of geodesic sphere, is a radial function. It is known that harmonic manifolds are Einstein [2]. They are naturally classified as per the sign of the Ricci constant. Let rr be the constant scalar curvature of MM.

  • If r=0r=0, then MM is flat, that is (M,g)=(n,Can)(M,g)=(\mathbb{R}^{n},Can) (Lemma 4.5).

  • If r>0r>0, then by Bonnet-Myer’s theorem MM is compact with finite fundamental group. They are compact rank one symmetric spaces by a well known result of Szabo (cf. [20]).

  • If r<0r<0, then MM is non-compact harmonic manifold. They are rank one symmetric spaces of non-compact type, if dimension of MM is atmost 55.

The main result in the theory of harmonic spaces is the Lichnerowicz Conjecture: Any simply connected, complete harmonic manifold is either flat or a rank one symmetric space. By the above classification, we see that the conjecture is resolved for compact harmonic manifolds and is open for non-compact harmonic manifolds of dimension 66. There are counter examples to the conjecture when dimension is atleast 77, known as the Damek-Ricci spaces or NA spaces. See for more details references in [20].

In the category of non-compact harmonic manifolds, we will be considering simply connected, complete, non-compact harmonic manifolds. It follows that, these spaces don’t have conjugate points (cf. [20]). Hence, by the Cartan-Hadamard theorem,

expp:TpMM\exp_{p}:T_{p}M\rightarrow M

is a diffeomorphism and every geodesic of MM is a line. That is, if γv:M\gamma_{v}:\mathbb{R}\rightarrow M is a geodesic of MM with vSpMv\in S_{p}M, γv(0)=v\gamma_{v}^{\prime}(0)=v, then d(γv(t),γv(s))=|ts|.d(\gamma_{v}(t),\gamma_{v}(s))=|t-s|.

Busemann function: Let γv\gamma_{v} be a geodesic line, then the two Busemann functions associated to γv\gamma_{v} are defined as [17]:

bv+(x)=limtd(x,γv(t))t,b_{v}^{+}(x)=\lim_{t\rightarrow\infty}d(x,\gamma_{v}(t))-t,
bv(x)=limtd(x,γv(t))t.b_{v}^{-}(x)=\lim_{t\rightarrow-\infty}d(x,\gamma_{v}(t))-t.

3. Einstein Solitons on Almost α\alpha-Cosymplectic 33-Manifolds

In this section, we examine the nature of a contact Einstein soliton on almost α\alpha-cosymplectic manifold. We also show that, the characteristic vector field ξ\xi is harmonic on almost α\alpha-cosymplectic 33-manifold admitting a contact Einstein soliton. Finally, we generalize Theorem 2.7 using these results.

Contact Einstein soliton: Let (M2n+1,g)(M^{2n+1},g) be a Riemannian manifold of dimension 2n+12n+1 (n1)(n\geq 1). Consider the Einstein soliton (2), with potential vector field VV, on an almost contact metric manifold (M,φ,ξ,η,g)(M,\varphi,\xi,\eta,g). Then the soliton is called contact Einstein soliton, if V=ξV=\xi that is, the potential vector field is the characteristic vector field.

The potential vector field VV is called transversal, if it is orthogonal to the characteristic vector field, that is VξV\perp\xi.

Theorem 3.1.

Let (M,φ,ξ,η,g)(M,\varphi,\xi,\eta,g) be an almost α\alpha-cosymplectic 33-manifold, admitting a contact Einstein soliton. Then the soliton is:

  1. (1)

    steady, if α2=σ2(a2+b2)(φe)(a)+e(b)\alpha^{2}=\sigma^{2}-(a^{2}+b^{2})-(\varphi e)(a)+e(b),

  2. (2)

    shrinking, if α2>σ2(a2+b2)(φe)(a)+e(b)\alpha^{2}>\sigma^{2}-(a^{2}+b^{2})-(\varphi e)(a)+e(b),

  3. (3)

    expanding, if α2<σ2(a2+b2)(φe)(a)+e(b)\alpha^{2}<\sigma^{2}-(a^{2}+b^{2})-(\varphi e)(a)+e(b).

Proof.

If the soliton is contact Einstein soliton, using V=ξV=\xi in (2), we have

g(Xξ,Y)+g(X,Yξ)+2g(X,QY)=(2λ+r)g(X,Y),g(\nabla_{X}\xi,Y)+g(X,\nabla_{Y}\xi)+2g(X,QY)=(2\lambda+r)g(X,Y), (10)

for any vector fields X,YX,Y on MM.
Substituting X=Y=ξX=Y=\xi in the above equation and using (8), we obtain

λ=2α22σ2r2.\lambda=-2\alpha^{2}-2\sigma^{2}-\frac{r}{2}. (11)

From the expression of rr (9), we get

λ=α2σ2+(a2+b2)+(φe)(a)e(b),\lambda=\alpha^{2}-\sigma^{2}+(a^{2}+b^{2})+(\varphi e)(a)-e(b), (12)

from which we can conclude the proof. ∎

Theorem 3.2.

Let (M,φ,ξ,η,g)(M,\varphi,\xi,\eta,g) be an almost α\alpha-cosymplectic 33-manifold, admitting a contact Einstein soliton. Then the characteristic vector field ξ\xi is harmonic.

Proof.

From (10), we get for X=ξX=\xi and Y=eY=e,

(φe)(σ)=2aσ.(\varphi e)(\sigma)=-2a\sigma. (13)

And for X=ξX=\xi and Y=φeY=\varphi e, from (10) we have

e(σ)=2bσ.e(\sigma)=2b\sigma. (14)

Now, using (13) and (14) in the expression of QξQ\xi in (8), we obtain

Qξ=(2α2+2σ2)ξ,Q\xi=-(2\alpha^{2}+2\sigma^{2})\xi,

which shows that ξ\xi is an eigenvector field of the Ricci operator QQ concluding the fact that ξ\xi is harmonic. ∎

We derive the identity involving the Lie derivative of the metric, Ricci operator, the potential vector field VV.

Lemma 3.3.

Let (M,g)(M,g) be a Riemannian manifold of scalar curvature rr, admitting an Einstein soliton (2). Then

𝔏Vg2=2dr(V)+4div((λ+r2)VQV),\|\mathfrak{L}_{V}g\|^{2}=2dr(V)+4\operatorname{div}\bigg{(}\bigg{(}\lambda+\frac{r}{2}\bigg{)}V-QV\bigg{)}, (15)

where QQ is the Ricci operator.

Proof.

In local coordinate system, (2) leads to

𝔏Vgij+Sij=(2λ+r)gij.\mathfrak{L}_{V}g^{ij}+S^{ij}=(2\lambda+r)g^{ij}.

Therefore,

𝔏Vg2=\displaystyle\|\mathfrak{L}_{V}g\|^{2}= Sij𝔏Vgij+(2λ+r)gij𝔏Vgij.\displaystyle-S^{ij}\mathfrak{L}_{V}g_{ij}+(2\lambda+r)g^{ij}\mathfrak{L}_{V}g_{ij}.
=\displaystyle= 𝔏Vr+gij𝔏VSij(2λ+r)gij𝔏Vgij.\displaystyle-\mathfrak{L}_{V}r+g_{ij}\mathfrak{L}_{V}S^{ij}-(2\lambda+r)g_{ij}\mathfrak{L}_{V}g^{ij}. (16)

Now,

gij𝔏VSij=\displaystyle g_{ij}\mathfrak{L}_{V}S^{ij}= gijVSijgijαViSαjgijαVjSiα\displaystyle g_{ij}\nabla_{V}S^{ij}-g_{ij}\nabla_{\alpha}V_{i}S^{\alpha j}-g_{ij}\nabla_{\alpha}V_{j}S^{i\alpha}
=\displaystyle= 2dr(V)2divQV.\displaystyle 2dr(V)-2\operatorname{div}QV. (17)

Observing that gij𝔏Vgij=2divVg_{ij}\mathfrak{L}_{V}g^{ij}=-2\operatorname{div}V and using (3) and (3), we get the required result. ∎

Now we derive the main result of this section.

Theorem 3.4.

Consider MM to be an almost α\alpha-cosymplectic 33-manifold, admitting a contact Einstein soliton. Then the following hold.

  1. (1)

    If σ0\sigma\neq 0, then α=a2+b22λ2+(φe)(a)e(b)\alpha=a^{2}+b^{2}-2\lambda^{2}+(\varphi e)(a)-e(b).

  2. (2)

    If σ=0\sigma=0, then MM is cosymplectic.

Proof.

Replacing XX by ee and YY by φe\varphi e, from (10) we get

g(eξ,φe)+g(e,φeξ)+2g(e,Qφe)=(2λ+r)g(e,φe).g(\nabla_{e}\xi,\varphi e)+g(e,\nabla_{\varphi e}\xi)+2g(e,Q\varphi e)=(2\lambda+r)g(e,\varphi e).

Using (7) and (8), after simplification we acquire,

ξ(σ)=σ2ασ.\xi(\sigma)=\sigma-2\alpha\sigma. (18)

Now putting X=e=YX=e=Y in (10) and using (7), (8), (9) and (12), we get

6α2+6σ24σμ+2α+r=0.6\alpha^{2}+6\sigma^{2}-4\sigma\mu+2\alpha+r=0. (19)

Similarly, putting X=φe=YX=\varphi e=Y in (10) and using (7), (8), (9) and (12), we also obtain

6α2+6σ2+4σμ+2α+r=0.6\alpha^{2}+6\sigma^{2}+4\sigma\mu+2\alpha+r=0. (20)

So comparing (19) and (20), we have σμ=0\sigma\mu=0. If σ0\sigma\neq 0, then from (20), we obtain the required result using (9).
Now suppose σ=0\sigma=0, then MM is α\alpha-cosymplectic. From [14], recall that an almost α\alpha-cosymplectic manifold MM is α\alpha-cosymplectic if and only if for any Xχ(M)X\in\chi(M),

QX=(α2+r2)X(3α2+r2)η(X)ξ.QX=\bigg{(}\alpha^{2}+\frac{r}{2}\bigg{)}X-\bigg{(}3\alpha^{2}+\frac{r}{2}\bigg{)}\eta(X)\xi. (21)

Since ξ\nabla\xi is symmetric, (10) becomes

g(Xξ,Y)+g(X,QY)=(λ+r2)g(X,Y).g(\nabla_{X}\xi,Y)+g(X,QY)=\bigg{(}\lambda+\frac{r}{2}\bigg{)}g(X,Y). (22)

Using (6) and (21), we have from (22), for any X,Yχ(M)X,Y\in\chi(M),

(α2+αλ)g(X,Y)(3α2+α+r2)η(X)η(Y)=0,(\alpha^{2}+\alpha-\lambda)g(X,Y)-\bigg{(}3\alpha^{2}+\alpha+\frac{r}{2}\bigg{)}\eta(X)\eta(Y)=0,

which implies α2+αλ=0\alpha^{2}+\alpha-\lambda=0 and 3α2+α+r2=03\alpha^{2}+\alpha+\frac{r}{2}=0.
That is λ=α2+α\lambda=\alpha^{2}+\alpha and r=6α22α=constantr=-6\alpha^{2}-2\alpha=\mbox{constant}, so that, λ+r2=2α2\lambda+\frac{r}{2}=-2\alpha^{2}.
Also, from (21), we have Qξ=2α2ξQ\xi=-2\alpha^{2}\xi which implies (λ+r2)ξQξ=0(\lambda+\frac{r}{2})\xi-Q\xi=0. Therefore, using Lemma 3.3 (15), we can say that ξ\xi is a Killing vector field, that is, ξ\nabla\xi is skew-symmetric. But in our case ξ\nabla\xi is symmetric, which implies ξ=0\nabla\xi=0, that is, α=0\alpha=0, proving the fact that MM is cosymplectic.

Corollary 3.5.

Consider MM to be a simply-connected, homogeneous, almost α\alpha-cosymplectic 33-manifold, admitting a contact Einstein soliton with σ=0\sigma=0. Then MM is an unimodular semidirect product Lie group GG of type G0μμ¯=2AG_{0\mu\overline{\mu}}=\mathbb{R}^{2}\rtimes_{A}\mathbb{R}, where A=(0μμ0)0A=\left(\begin{array}[]{cc}0&\mu\\ -\mu&0\end{array}\right)\neq 0, is a real matrix. Moreover, GG is the Lie group E~2\tilde{E}^{2} equipped with its flat left invariant cosymplectic structure.

Proof.

The proof follows from Theorem 2.7 and Theorem 3.4. ∎

4. Ricci Solitons on Harmonic Manifolds

Recall that the Ricci solitons are solutions of (1). Clearly, if a manifold is Einstein of constant rr, then trivial solitons X=0X=0 and XX a Killing vector field are solutions of (1) with λ=r\lambda=r.

In this section, we study Ricci solitons on complete, simply connected, harmonic manifolds. We prove a Lichnerowicz type result that, a harmonic manifold admits a non-trivial Ricci soliton if and only if MM is flat. More precisely, we show that compact harmonic manifolds and non-flat harmonic manifolds do not admit non-trivial Ricci solitons. But flat harmonic manifolds do admit non-trivial shrinking and expanding Ricci solitons.

In the sequel, harmonic manifold means complete, simply connected harmonic manifold.

The main theorem of this section is:

Theorem 4.1.

Let (M,g)(M,g) be a harmonic manifold. Then MM admits a non-trivial Ricci soliton if and only if MM is flat. In this case, the steady Ricci soliton is trivial of Killing type given by X=bv;X=\nabla{b_{v}}^{-}; where bv(x)=x,vb_{v}^{-}(x)=-\langle x,v\rangle, the Busemann function, is the potential function on MM. In case, the Ricci soliton is shrinking or expanding, the potential function is given by f(x)=λd(x,p)2+f(p)f(x)=\lambda{d(x,p)}^{2}+f(p), for constant λ0\lambda\neq 0; and point pp is the minimum or the maximum of ff and X=fX=\nabla f is the corresponding non-trivial Ricci soliton.

Corollary 4.2.

There are no deformations of harmonic manifolds, and in particular, of rank one symmetric spaces under a Ricci soliton. In particular, we obtain a strengthening of [1], and also the stability of the compact rank one symmetric spaces under a Ricci soliton.

4.1. Proof of Theorem 4.1

In this subsection we prove Theorem 4.1. We begin with the following important proposition.

Proposition 4.3.

If a complete manifold admits a Ricci soliton, then it is a gradient soliton.

Proof.

This follows from Remark 3.2 of Perelman [13] (see also page 22 of [12])). Hence, in this case, if XX is a Ricci soliton, then X=fX=\nabla f, for some smooth function f:MRf:M\rightarrow R. ∎

Remark 4.4.

Here we are only concerned with simply connected and complete Riemannian manifold. In this case, clearly, we can write X=fX=\nabla f by Poincaré Lemma, for some fC(M)f\in C^{\infty}(M).

Lemma 4.5.

Ricci flat harmonic manifold is flat.

Proof.

It can be shown that any harmonic manifold (M,g)(M,g) is asymptotically harmonic [20]. That is there exists a constant h0h\geq 0 such that

Δbv±=h.\Delta{b_{v}}^{\pm}=h.

Let Lt=2bv+L_{t}={\nabla}^{2}{b_{v}}^{+} denote the second fundamental form of horosphere, bv1(t)b_{v}^{-1}(t). Then LtL_{t} satisfies the Riccati equation, that is for xt{γ(t)}x_{t}\in\{\gamma^{\prime}(t)\}^{\perp},

Lt(xt)+Lt2(xt)+R(xt,γ(t))γ(t)=0.{L_{t}}^{\prime}(x_{t})+{L_{t}}^{2}(x_{t})+R(x_{t},\gamma^{\prime}(t))\gamma^{\prime}(t)=0.

Tracing the above equation, we obtain that trLt2=0\operatorname{tr}{L_{t}}^{2}=0, as Ricci(γ(t),γ(t))=0(\gamma^{\prime}(t),\gamma^{\prime}(t))=0. But as LtL_{t} is a symmetric operator on {γ(t)}\{\gamma^{\prime}(t)\}^{\perp}, Lt=0L_{t}=0. Consequently, R(x,v)v=0R(x,v)v=0 for any xvx\in{v}^{\perp} and for any vSMv\in SM. Thus (M,g)(M,g) is flat. ∎

Proposition 4.6.

If a harmonic manifold admits a Ricci soliton, then it admits a Gaussian.

Proof.

As in this case (M,g)(M,g) is Einstein, it follows that

2f=2(λr)g,\nabla^{2}f=2(\lambda-r)g, (23)

where rr is a constant scalar curvature of MM. Thus ff is a Gaussian, that is it satisifes (23). ∎

Lemma 4.7.

Let X=fX=\nabla f be a Killing vector field on compact harmonic manifold, then XX is trivial. Trivial solitons of Killing type do not exist on non-compact, non-flat harmonic manifold. On flat harmonic manifold, Killing vector field is X=bvX=\nabla{b_{v}}^{-}, where bv(x)=x,vb_{v}^{-}(x)=-\langle x,v\rangle is a Busemann function on n\mathbb{R}^{n}.

Proof.

Because X=fX=\nabla f is a non-trivial Killing vector field, we have

2f=0.\nabla^{2}f=0.

Therefore, f=constant0\rVert\nabla f\rVert=\mbox{constant}\neq 0, consequently, ff has no critical points. Any Killing vector field of constant norm satisfies (p. 164-167, [17]):

2f2=Ric(f,f).\displaystyle{\rVert\nabla^{2}f\rVert}^{2}=\mbox{Ric}(\nabla f,\nabla f).

Therefore,

0=2f=rf2\displaystyle 0=\rVert\nabla^{2}f\rVert=r{\rVert\nabla f\rVert}^{2}

This implies that for ff non-constant, r=0r=0 and therefore Ric0\mbox{Ric}\equiv 0 and hence harmonic manifold must be flat (Lemma 4.5).
We have f=constant\rVert\nabla f\rVert=\mbox{constant}. We may assume that f=1\rVert\nabla f\rVert=1, therefore ff is distance function which is harmonic function on (n,Can)({\mathbb{R}}^{n},Can). By Proposition 5.1 of [20], it follows that

f(x)=bv(x)=x,v,f(x)=b_{v}^{-}(x)=-\langle x,v\rangle,

is a Busemann function on n{\mathbb{R}}^{n} [17].
If MM is compact, 2f=0\nabla^{2}f=0 implies that ff is a harmonic function. Hence, ff must be a constant function. ∎

Proposition 4.8.

A compact harmonic manifold (M,g)(M,g) does not admit a non-trivial Ricci soliton.

Proof.

We have,

2f=2(λr)g.{\nabla}^{2}f=2(\lambda-r)g.

Therefore, Δf=2(λr)n\Delta f=2(\lambda-r)n implies by the Bochner’s formula that,

12Δ(f2)=4(λr)2n2+r(f2).\displaystyle\frac{1}{2}\Delta(\|\nabla f\|^{2})=4(\lambda-r)^{2}n^{2}+r(\|\nabla f\|^{2}). (24)

Therefore,

4(λr)2n2Vol(M)=rMf2<0.4(\lambda-r)^{2}n^{2}\operatorname{Vol}(M)=-r\int_{M}\|\nabla f\|^{2}<0.

This implies that f=0\|\nabla f\|=0, therefore ff is constant. ∎

Lemma 4.9.

A non-compact, harmonic manifold admits a non-trivial Ricci soliton if and only if it is flat. The flat harmonic manifold admits shrinking and expanding Ricci solitons with the corresponding potential function, f(x)=λd(p,x)2+f(p)f(x)=\lambda{d(p,x)}^{2}+f(p), for some pMp\in M.

Proof.

Suppose that a non-compact, harmonic manifold admits a non-trivial Ricci soliton. Therefore, it admits a Gaussian with (λr)0(\lambda-r)\neq 0.

2f=2(λr)g.{\nabla}^{2}f=2(\lambda-r)g.

Therefore, ff is either convex or concave function. Consequently, the only possible critical point of ff is either maximum or minimum of ff. Suppose that pp is a critical point of ff. Note that along any unit speed geodesic of MM starting from pp,

f′′(t)=2(λr).\displaystyle f^{\prime\prime}(t)=2(\lambda-r). (25)

Therefore, f(t)=2(λr)t+cf^{\prime}(t)=2(\lambda-r)t+c. Hence, there is exactly one critical point, and hence c=0c=0. Thus, f(t)=(λr)t2+f(p),f(t)=(\lambda-r)t^{2}+f(p), consequently ff is a radial function. This implies that,

Δf=f′′+ΘΘf=2(λr)n.\Delta f=f^{\prime\prime}+\frac{{\Theta}^{\prime}}{\Theta}f^{\prime}=2(\lambda-r)n.

Therefore,

f′′+ΘΘ2(λr)t=2(λr)n.f^{\prime\prime}+\frac{{\Theta}^{\prime}}{\Theta}2(\lambda-r)t=2(\lambda-r)n.

Consequently by (25),

Θ(t)Θ(t)=n1t.\frac{{\Theta}^{\prime}(t)}{\Theta(t)}=\frac{n-1}{t}.

Comparing with the series expansion (see (4.4) of [20]),

Θ(t)Θ(t)=n1trt3+,\frac{{\Theta}^{\prime}(t)}{\Theta(t)}=\frac{n-1}{t}-\frac{r\;t}{3}+\cdots,

we obtain r=0r=0, hence MM is flat. Finally, f(x)=λd(p,x)2+f(p)f(x)=\lambda{d(p,x)}^{2}+f(p) follows from section 11 of [5]. ∎

Finally we come to the proof of Theorem 4.1.

Proof: A compact harmonic manifold can’t admit non-trivial Ricci soliton (Proposition 4.8). If a non-compact harmonic manifold admits a trival Ricci soliton of Killing type, then (λr)=0(\lambda-r)=0, implies that r=0.r=0. Therefore, MM is flat and X=bvX=\nabla{b_{v}}^{-} (Lemma 4.7). If a non-compact harmonic manifold admits a non-trival Ricci soliton, then (λr)0(\lambda-r)\neq 0 again implies that r=0,r=0, and MM is flat. In this case X=fX=\nabla f, where f(x)=λd(p,x)2+f(p)f(x)=\lambda{d(p,x)}^{2}+f(p), for some pMp\in M (Lemma 4.9).

Remark 4.10.

We have shown that Theorem 4.1 confirms Theorem 2.1 in case of harmonic manifolds. Also Theorem 4.1 implies that there are no non-trivial deformation of non-flat harmonic manifolds. This indicates a result supporting the conjecture that, there are no non-trivial deformations of harmonic manifolds; and hence there should be only finitely many classes of harmonic manifolds.

5. Acknowledgements

Dr. Naeem Ahmad Pundeer would like to thank to U.G.C. for its Dr. D.S. Kothari Postdoctoral Fellowship. The corresponding author, Mr. Paritosh Ghosh, thanks UGC Junior Research Fellowship of India. The authors also would like to thank Mr. Dipen Ganguly for his wishful help in this research.

References

  • [1] Balmer, R., Stability of symmetric spaces of noncompact type under Ricci flow, Geom. Funct. Anal., 25 (2015), 342-416.
  • [2] Besse, A.L., Manifolds all of whose geodesics are closed, Springer-Verlag, Berlin Heidelberg, (1978).
  • [3] Blair, D.E., Riemannian geometry of contact and symplectic manifolds, Progress in Mathematics, Birkhäuser, New York, (2010).
  • [4] Catino, G. and Mazzieri, L., Gradient Einstein solitons, Nonlinear Anal., 132 (2016), 66–94.
  • [5] Cheeger, J. and Colding, T., Lower bounds on Ricci curvature and the almost rigidity of warped products, Ann. Math., 144 (1996), 189-237.
  • [6] Cunha, A.W. and Griffin, E., On non-compact gradient solitons, arXiv:2207.05822, (2022).
  • [7] Eells, J. and Sampson, J.H., Harmonic Mappings of Riemannian Manifolds, Amer. J. Math., 86 (1964), 109-160.
  • [8] Hamilton, R.S., Three manifolds with positive Ricci curvature, J. Diff. Geom., 17 (1982), 255-306.
  • [9] Hu, Q., Xu, G. and Yu, C, The rigidity and stability of gradient estimates, J. Geom. Anal., 32 (2022), 1-13.
  • [10] Li, J. and Liu, X., Ricci solitons on homogeneous almost α\alpha-cosymplectic three-Manifolds, Mediterr. J. Math., 19 (2022), 1-12.
  • [11] Libermann, P., Sur les automorphismes infinitésimaux des structures symplectiques et des structures de contact, Colloque Géom. Diff. Globale, (1959), 37–59.
  • [12] Mondal, C. K. and Shaikh, A. A., Some results in η\eta-Ricci soliton and gradient ρ\rho-Einstein soliton in a complete Riemannian manifold, Commun. Korean Math. Soc., 34 (2019), 1279-1287.
  • [13] Perelman, G., The entropy formula for the Ricci flow and its geometric applications, arXiv:math  211159, (2002).
  • [14] Perrone, D., Classification of homogeneous almost α\alpha-coKa¨\ddot{a}hler three-manifolds, Diff. Geom. Appl., 59 (2018), 66–90.
  • [15] Perrone, D., Classification of homogeneous almost cosymplectic three manifolds, Diff. Geom. Appl., 30 (2012), 49–58.
  • [16] Perrone, D., Left-invariant almost α\alpha-coKähler structures on 3D3D semidirect product Lie groups, Int. J. Geom. Meth. Mod. Phys., 16 (2019), 1-18.
  • [17] Petersen, P., Riemannian geometry, New York, Springer-Verlag, (2006).
  • [18] Petersen, P. and Wylie, W., Rigidity of gradient Ricci solitons, Pac. J. Math., 241 (2009), 329-345.
  • [19] Ranjan, A. and Shah, H., Harmonic manifolds with minimal horospheres, J. Geom. Anal., 12 (2002), 683-694.
  • [20] Ranjan, A. and Shah, H., Busemann functions in a harmonic manifold, Geom. Dedicata, 101 (2003), 167-183.
  • [21] Tashiro, Y., Complete Riemannian manifolds and some vector fields, Trans. Amer. Math. Soc., 117 (1965), 251-275.